Nanotribology of Ionic Liquids as Lubricant Additives for Alumina

Dec 5, 2017 - School of Molecular Sciences, University of Western Australia, Perth, WA 6009, Australia. J. Phys. Chem. C , 2017, 121 (51), pp 28348–...
0 downloads 0 Views 602KB Size
Subscriber access provided by READING UNIV

Article

Nanotribology of Ionic liquids as Lubricant Additives for Alumina Surfaces Stephen Cowie, Peter K Cooper, Rob Atkin, and Hua Li J. Phys. Chem. C, Just Accepted Manuscript • DOI: 10.1021/acs.jpcc.7b09879 • Publication Date (Web): 05 Dec 2017 Downloaded from http://pubs.acs.org on December 6, 2017

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

The Journal of Physical Chemistry C is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 19 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Nanotribology of Ionic liquids as Lubricant Additives for Alumina Surfaces Stephen Cowie,1,# Peter K. Cooper,2,# Rob Atkin2,* and Hua Li2,* 1

Priority Research Centre for Advanced Fluids and Interfaces, University of Newcastle, Callaghan, NSW 2308, Australia 2 School of Molecular Sciences, University of Western Australia, WA 6009, Australia #

Both authors contributed equally to this work.

Corresponding author: [email protected]; [email protected]

Abstract Common lubricants, optimized for steel, perform poorly for alumina surfaces, making ionic liquids (ILs) attractive potential alternatives, either in neat form or as oil additives. Here, atomic force microscopy (AFM) has been used to study the lubricity of two oil – miscible ILs between a silicon AFM tip and an alumina surface. Trihexyltetradecylphosphonium bis(2,4,4trimethylpentyl)phosphinate (P6,6,6,14 (iC8)2PO2) and trihexyltetradecylphosphonium bis(2ethylhexyl) phosphate (P6,6,6,14 DEHP) were mixed with hexadecane at concentrations from 0 to 100 mol% IL. Both ILs are effective lubricants for alumina, reducing lateral forces to approximately a third of the forces measured in hexadecane. When used as an additive in hexadecane, increasing the IL concentration generally decreased the friction, as the IL adsorbs on alumina and forms a strong boundary layer. P6,6,6,14 DEHP mixtures reduce friction more effectively than P6,6,6,14 (iC8)2PO2. For both ILs, a 2 mol% mixture of IL and hexadecane reduced friction most effectively, even more than the neat IL.

1 ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 2 of 19

Introduction Lubricants reduce friction, adhesion and wear between sliding surfaces. Under low loads, viscous hydrodynamic forces separate the surfaces and prevent physical contact.1 At high loads, the bulk of the lubricant is squeezed out, and only a single boundary layer of lubricant molecules remains between the sliding surfaces.2 Common hydrocarbon-based oils adsorb weakly to surfaces, so oil-miscible surfactants and polymers are added which adsorb to the surface and remain in place even under high loads.1-2 The effectiveness of surfactant and polymer additives depends on the affinity it has for the surface via Coulombic, van der Waals, H-bonding, and solvophobic interactions. Unfortunately, many common additives which are effective on steel surfaces are not effective on light-weight surfaces such as aluminum and aluminum alloys.3-4 Thus, there is a push towards developing additives to match these surfaces. Ionic liquids (ILs) are pure salts that have melting points below 100 °C.5 ILs are high performing lubricants in neat (undiluted) form for a variety of surfaces, including silica, steel, aluminum, titanium at both the macro-6-12 and nano-13-17 scales. Like conventional high-end lubricants, such as perfluoropolyethers, ILs have high thermal stability and low vapor pressure, enabling them to operate in extreme conditions, e.g. high temperature and low pressure. Additionally, IL ions interact strongly with surfaces and each other via Coulombic, van der Waals, H-bonding and solvophobic interactions18-19; this means they resist being ‘squeezed-out’, meaning a lubricating film remains in place up to higher loads than for a comparable conventional lubricant.14, 16-17, 20-25 As the strength of IL boundary layer is related to chemical composition and structure (i.e. alkyl chain length and charge localization), control of IL lubricity can be achieved by tuning the composition and orientation of ions in the boundary layer by changing IL molecular structure and/or surface properties. 2 ACS Paragon Plus Environment

Page 3 of 19 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

ILs are relatively expensive compared with current commercially available lubricants. However, the recent development of oil miscible ILs means they can diluted in cheaper base oils and used as additives.26-34 IL/oil mixtures investigated at the macroscale outperform conventional additives at comparable concentrations,26-33 but the lubrication mechanisms of using ILs as additives remain only partially understood.27 Nanotribology studies are less affected by surface asperities and roughness than macroscale measurements, and thus are able to reveal lubrication mechanisms.21 Recent macro- and nanotribology studies have demonstrated the efficiency of two oil miscible ILs, trihexyl(tetradecyl)phosphonium bis(2,4,4-trimethylpentyl)phosphinate (P6,6,6,14 (iC8)2PO2) and trihexyltetradecylphosphonium bis(2-ethylhexyl) phosphate (P6,6,6,14 DEHP) as neat lubricants and as lubricant base oil additives.11, 27 Nanotribology experiments of P6,6,6,14 (iC8)2PO2/hexadecane mixtures on silica and titanium surfaces reveal that a minimum concentration of IL (~ 1 – 2 mol%) is required to form a robust boundary layer. Due to its high strength to weight ratio, aluminum and its alloys have been widely applied in capacitors, lightweight mechanical components, and optical/analytical devices.35-37 Furthermore, aluminum forms a passivating oxide layer in air that protects the bulk aluminum from further corrosion, unlike iron, which corrodes destructively.35-37 One of the drawbacks of aluminum is that common lubricants designed for steel surfaces cannot lubricate alumina surfaces effectively, leading to high friction and wear.9, 27, 38 The alumina surface is net positively charged in water39 at neutral pH, and based on past experience,40 a net positive charge is expected to persist in IL/oil mixtures. This contrasts with previously investigated net negatively charged silica and titania surfaces,40-41 and thus the adsorbed boundary layer on alumina is expected to be anion rich rather than cation rich like for silica and titania. Macrotribology investigations for ILs as neat lubricants8,

11-12

and IL/oil 3

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 19

mixtures as lubricant additives27, 34 on aluminum-steel interfaces have shown the effectiveness of these ILs. However, the lubrication mechanisms are currently poorly understood. No fundamental nanotribology studies have been conducted on alumina surfaces, especially in the boundary layer lubrication regime, where the bulk lubricant film has been squeezed out and only remains a surface adsorbed boundary layer. The effects of ion structure and concentration on the adsorbed boundary layer and lubricity are also unclear. In this work, we investigate the nanotribology of two oil miscible ILs, P6,6,6,14 i(C8)2PO2 and P6,6,6,14

DEHP, as lubricant additives for alumina surfaces in the boundary layer lubrication

regime. The outcomes reveal the relationship between IL structure and boundary layer lubrication, and thus help to elucidate the lubrication mechanisms of IL/oil mixtures and aid the development of new, cost-effective lubricant additives for aluminum and aluminum alloy systems.

Materials and Methods Flat alumina surfaces (10 mm × 10 mm × 0.5 mm, purity 99.6%) were purchased from MTI Corporation, CA. The morphology of the alumina surface was investigated using contact mode AFM with a scan size of 1 µm × 1 µm, cf. Figure S1; the surface roughness is 4±1 nm. Hexadecane is used as a model apolar lubricant base oil due to its simple structure, well-known properties, and commercial availability in high purity. Hexadecane (ReagentPlus®, 99%) was purchased from Sigma-Aldrich. Both ILs, P6,6,6,14 DEHP (purity > 98%) and P6,6,6,14 i(C8)2PO2 (purity > 95%), provided by Iolitec, are chemically stable in air and fully miscible in non-polar base oils. The zero shear viscosities of the IL/hexadecane mixtures at 20 °C were measured by an AR-G2 rheometer (TA instruments), and are provided in Table S1 and Table S2.

4 ACS Paragon Plus Environment

Page 5 of 19 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

All nanotribological measurements were determined using a Bruker Nanoscope MultiMode 8 AFM with an EV scanner in contact mode. AFM cantilevers (spring constant = 2.0 ± 0.2 N/m by the thermal tune method, tip radius ~8 nm) from the same batch (model NSC36 with Al backside coating, Mikromasch, Tallinn, Estonia) were used over the course of the investigation. The cantilevers were cleaned with ethanol and Milli-Q water (resistivity 18.2 MΩ·cm at 25 °C) dried under nitrogen and irradiated with UV light for at least 15 min before use. The sample volume of the AFM liquid cell was 0.1 mL. 0.5 mL of the sample was flushed in the AFM liquid cell before equilibrating for 20 min. Friction (lateral) forces were obtained by performing AFM scans with a scan angle of 90° (with respect to the long axis of the cantilever) and with the slow scan axis disabled.42 The scan size was 100 nm, and scan rate was 30 Hz. The lateral deflection signal (i.e., cantilever twist) was converted to lateral force using a customized function produced in Matlab R2015a which takes into account the torsional spring constant and the geometrical dimensions of the cantilever. These experiments were repeated more than three times; the resulting lateral force data showed the same trend with small variations. Examples of repeated results for 0.5 mol% P6,6,6,14 (iC8)2PO2/hexadecane and 0.5 mol% P6,6,6,14 DEHP/hexadecane mixtures are shown in Figure S2.

Results and Discussion The nanotribology of the IL, hexadecane and various IL/hexadecane concentrations was investigated with AFM. The lateral (friction) force as a function of normal load up to 300 nN is presented in Figure 1 and Figure 2. Loads in mechanical engine parts and disc brakes generally do not exceed 0.5 GPa, which is corresponds to a load of ~70 nN in this study.41, 43 Loads up to 300 nN therefore probe IL lubrication for extreme conditions. Friction coefficients (µ) were calculated using a modified version of Amontons’ Law, FL = FL(0) + µFN, where FL(0) is the 5 ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 19

lateral force at zero applied normal load, dependent on the adhesion, roughness and compressibility of both the AFM tip and surface. 44 Hexadecane The lateral force data for the AFM tip sliding over the alumina surface in hexadecane is similar to previous studies on weakly absorbing liquids at various surfaces, including silica and titania 41, 44-46

. At zero normal load there is a small lateral force (~10 nN) caused by adhesion between the

silica tip and the alumina surface47, consistent with the slightly attractive forces detected in the normal force-separation profiles, cf. Figure S3(a). Above 10 nN the lateral force increases at a constant rate, giving a friction coefficient of µ = 1.0 ± 0.3. An average data set for hexadecane on alumina is shown in Figure 1 and Figure 2; however, some variability (±90 nN for lateral force measurements at a normal load of 300 N) was observed for hexadecane which was not observed for the IL/hexadecane mixtures. The variability is likely due to variation in surface roughness and/or humidity for different tests. As expected, hexadecane is a poor lubricant and shows high friction because it has weak interaction with alumina and thus adsorbs weakly to the alumina surface; therefore hexadecane is easily pushed through by the AFM tip even at low normal loads, resulting in direct contact of the AFM tip and the alumina surface. P6,6,6,14 (iC8)2PO2 Results for lateral force vs normal load data for P6,6,6,14 (iC8)2PO2 on alumina reveal excellent lubrication, similar to previous studies on silica and titania.41, 44, 46 Unlike in hexadecane, at zero normal load negligible lateral force is measured in P6,6,6,14 (iC8)2PO2, in line with negligible attraction detected in the normal force-separation data provided in Figure S3(b). The lateral forces increase with normal load at a constant rate (µ = 0.25 ± 0.01) which is around three times lower than that of hexadecane. The much lower lateral force and friction coefficient for P6,6,6,14 6 ACS Paragon Plus Environment

Page 7 of 19 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

(iC8)2PO2 is consistent with the formation of an IL boundary layer formed by P6,6,6,14 (iC8)2PO2 ions adsorbed to the alumina surface.11, 27 However, at a normal load of ~230 nN, lateral forces of P6,6,6,14 (iC8)2PO2 begin to increase at a faster rate (µ = 0.55 ± 0.04). This suggests a change takes place in the boundary layer between the AFM tip and the alumina surface at high loads. The change is consistent with the rupture of the lubricating boundary layer by the AFM tip at high loads, as proposed previously for the same IL on titania surfaces.41 Interestingly, although the surface roughness of alumina is higher than that of silica and titania used in previous studies, the friction coefficient of P6,6,6,14 (iC8)2PO2 on alumina (µ = 0.25 ± 0.01) before the breakpoint is slightly lower than that on silica (µ = 0.34 ± 0.03) and titania (µ = 0.40 ± 0.04). Silica and titania are negatively charged surfaces, therefore the IL boundary layer on these surfaces is expected to be cation-rich. Alumina, in contrast, is net positively charged, and therefore the IL boundary layer is anion-rich. This indicates that the (iC8)2PO2- anions at positively charged alumina surfaces are more lubricating than the P6,6,6,14+ cations at negatively charged silica and titania surfaces. Lateral forces for 0.1 mol% P6,6,6,14 (iC8)2PO2 in hexadecane (µ =0.95 ± 0.04) are similar to hexadecane (µ =1.0 ± 0.3) with reduced variability. An earlier study showed that the IL surface excess increases with the bulk IL concentration.46 At 0.1 mol%, the P6,6,6,14 (iC8)2PO2 surface excess at the alumina interface is too low to form an effective boundary layer; the silica tip likely slides directly on the surface as it does in hexadecane, resulting in similar lateral force data to hexadecane. The improved reproducibility indicates the added IL assists to reduce the variation in surface roughness and adhesive forces. When the P6,6,6,14 (iC8)2PO2 concentration is increased to 0.5 mol%, at low normal loads (< 130 nN), the lateral force increases slowly and is comparable to the neat P6,6,6,14 (iC8)2PO2. However,

7 ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 19

when the normal load reaches 130 nN, the lateral force increases sharply, until at very high loads (> 250 nN) the lateral force increases at a similar rate to neat hexadecane and the lower concentration IL/oil mixtures. These data indicate that while ions are adsorbed to the surface at this IL concentration, the adsorbed ion density is relatively low, resulting in a poorly formed boundary layer. Thus the breakpoint, the critical normal load at which the lateral force increases rapidly, is lower than neat P6,6,6,14 (iC8)2PO2. At higher IL concentrations (≥ 1 mol%), the friction coefficients before breakpoints are similar to that of the neat P6,6,6,14 (iC8)2PO2, cf. Table 1 and Figure 3. The breakpoint for 1 mol% P6,6,6,14 (iC8)2PO2/hexadecane mixture occurs at 170 nN, while for 5 mol% and neat P6,6,6,14 (iC8)2PO2 there is a breakpoint at approximately 230 nN. Generally the nanofriction data for P6,6,6,14 (iC8)2PO2/hexadecane mixtures on alumina shows similar trends with those obtained on silica and titania.41, 44, 46 At low bulk IL concentrations (≤ 1.0 mol%) the lower surface excess of ions leads to a lower density of ions in the boundary layer which the tip displaces at high force. At higher IL concentrations, the higher surface excess leads to a more robust boundary layer, and the lateral forces are similar to that of the neat IL. Thus, in general, lateral forces decrease and the breakpoint increases with IL concentration. A clear exception to this trend is 2 mol% P6,6,6,14 (iC8)2PO2, with lower later forces than all other P6,6,6,14 (iC8)2PO2/hexadecane concentrations and no obvious breakpoint. The reason for this will be discussed later in the manuscript.

8 ACS Paragon Plus Environment

Page 9 of 19 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Figure 1. Lateral Force vs normal load in P6,6,6,14 (iC8)2PO2/hexadecane mixtures for an AFM tip (r ≈ 8 nm) sliding on an alumina surface at 6 µm s-1. The structure of P6,6,6,14 (iC8)2PO2 is shown in the inset.

Table 1. Friction coefficients (µ) for P6,6,6,14 (iC8)2PO2/hexadecane mixtures before and after breakpoints. mol% P6,6,6,14

Hexadecane

0.10%

0.50%

1%

2%

5%

(iC8)2PO2 µ (before break)

(iC8)2PO2 1.0 ± 0.3

0.95 ± 0.04

µ (after break) Breakpoint (nN)

P6,6,6,14

n/a

n/a

0.30 ± 0.02

0.28 ± 0.02

0.76 ± 0.09

0.61 ± 0.04

130

170

0.26 ± 0.02

n/a

0.28 ± 0.01

0.25 ± 0 .01

0.74 ± 0.09

0.55 ± 0.04

230

230

P6,6,6,14 DEHP Friction measured in P6,6,6,14 DEHP is very low (µ = 0.21 ± 0.02). At high loads (~230 nN), there is a sharp increase in the lateral force (Figure 2), similar to what is seen for P6,6,6,14 (iC8)2PO2. At low to medium loads, an adsorbed IL boundary layer between the surfaces prevents direct

9 ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 19

contact; however, at high loads, the AFM tip pushes through the boundary layer and it cannot reform within the time frame of the experiment, resulting in increased friction. At low concentrations (< 0.1 mol%), the friction of P6,6,6,14 DEHP/hexadecane mixtures is higher than that of neat P6,6,6,14 DEHP but lower than that of hexadecane. In P6,6,6,14 DEHP/hexadecane mixtures there are no obvious breakpoints up to 300 nN. Furthermore, when the concentration of P6,6,6,14 DEHP is 0.1 mol% and higher, the lateral forces and friction coefficients are significantly reduced and comparable to those of neat P6,6,6,14 DEHP before the breakpoint, cf. Figure 2 and Figure 3. These results indicate the formation of robust IL ion boundary layers at such IL concentrations. Interestingly, as observed for the P6,6,6,14 (iC8)2PO2/hexadecane mixtures, 2 mol% P6,6,6,14 DEHP (µ = 0.09) exhibits superior tribological performance over any other solution examined, even neat P6,6,6,14 DEHP (µ = 0.21). Additionally, we have found 95 mol% P6,6,6,14 DEHP (Figure S4) has no observable breakpoints up to a normal load of 300 nN. These data indicate that the integration of hexadecane in the boundary layer plays a vital part in the improvement of lubricity. Although the IL and solvent are different, these results are in line with Seddon et al’s findings on molecular ordering of diluted ILs at the interface, which show that a small amount of either solvent or solute could increase the strength of boundary layers.48

10 ACS Paragon Plus Environment

Page 11 of 19 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Figure 2. Lateral force vs normal load in P6,6,6,14 DEHP/hexadecane mixtures for an AFM tip (r ≈ 8 nm) sliding on an alumina surface at 6 µm s-1. The structure of P6,6,6,14 DEHP is shown in the inset.

Table 2. Friction coefficients (µ) of P6,6,6,14 DEHP/Hexadecane mixtures on alumina. mol% DEHP

µ

0.01%

0.03%

0.10%

0.50%

1%

2%

5%

IL Breakpoint: 240 nN before break point

after break point

0.45

0.42

0.24

0.15

0.13

0.09

0.25

0.21

1.0

±0.02

±0.02

±0.02

±0.01

±0.01

±0.01

±0.02

±0.02

±0.1

11 ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 19

Figure 3. Friction coefficients of P6,6,6,14 (iC8)2PO2/hexadecane and P6,6,6,14 DEHP/hexadecane mixtures on alumina before the break point.

Comparison of P6,6,6,14 (iC8)2PO2 and P6,6,6,14 DEHP ILs reduce friction by adsorbing to the solid surface and providing a low shear boundary layer which prevents contact between surfaces. At high loads, this boundary layer can be ruptured, resulting in direct contact between the surfaces, and a sharp increase in friction as seen in Figure 1 and Figure 2. The load at which the boundary layer is ruptured depends on: (i) the strength of the interaction between the IL ions in the boundary layer and the surface, (ii) the surface excess of IL at the interface, and (iii) the mobility of ions into the boundary layer. The molecular structure of the IL determines (i) the strength of the interaction between the IL ions in the boundary layer and the surface. Furthermore, we have shown in previous studies46 as well as this study that (ii) the surface excess of IL at the interface depends on the bulk IL concentration. Until now, the importance of (iii) the mobility of the ions into the boundary layer is not clear. Both ILs in this study are highly viscous (> 1000 mPa·s) at room temperature,49 which means that at high loads when the tip displaces the boundary layer IL ions, the ability of the ions to replenish the boundary layer is limited. As hexadecane is significantly less viscous than the ILs 12 ACS Paragon Plus Environment

Page 13 of 19 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

investigated (3 mPa·s vs ~1000 mPa·s), inclusion of hexadecane within the boundary layer may increase the mobility of ions into the boundary layer, and thus once disrupted, the boundary layer is able to replenish immediately, similar to measurements in aqueous salt solutions.50-51 This is likely why neat P6,6,6,14 DEHP presents an obvious breakpoint at a normal load of 240 nN, whereas 0.1 mol% to 95 mol% P6,6,6,14 DEHP/hexadecane mixtures do not (Figure 2 and Figure S4). However, the presence of hexadecane in the boundary layer may also reduce the packing and ordering of ions, thus reducing the resistance of the boundary layer under high normal loads. 2 mol% IL/hexadecane mixtures show the best lubricity because the boundary layer at this concentration is robust enough and replenishes quickly under high normal loads. In the neat form, P6,6,6,14 DEHP shows lubricity comparable to P6,6,6,14 (iC8)2PO2 on alumina. However, when mixed with hexadecane, P6,6,6,14 DEHP is more effective than P6,6,6,14 (iC8)2PO2; the critical concentration for P6,6,6,14 DEHP/hexadecane (0.1 mol%) is much lower than that of P6,6,6,14 (iC8)2PO2 (1.0 mol%), as well as the friction coefficients of the same concentration, cf. Figure 3. This suggests that when mixed with hexadecane, P6,6,6,14 DEHP adsorbs more effectively to a charged alumina surface than P6,6,6,14 (iC8)2PO2, thereby creating a smoother and more robust ion boundary layer, leading to reduced friction. ILs have been reported to be only partially dissociated, with some neutral ion-pairs or clusters.52 Ion diffusion rates for different structured ILs vary, but are generally proportional to the electrical conductivity. The conductivities of P6,6,6,14 DEHP and P6,6,6,14 (iC8)2PO2 are the same, both are 3.6 µS/cm.27 This means the diffusion rates of the two ILs in the neat form are comparable, which is consistent with the similar friction coefficients found in the neat form. However, the conductivity of P6,6,6,14 DEHP/oil mixture is about two orders of magnitude higher than that of P6,6,6,14 (iC8)2PO2/oil mixture.27 This suggests that at the same IL concentration, a larger portion of P6,6,6,14 DEHP

13 ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 19

dissociates to single ions than P6,6,6,14 (iC8)2PO2 when dissolved in apolar oils, leading to higher ion adsorption at the interface and more robust boundary layers and thus low friction. The nanotribology results for P6,6,6,14 (iC8)2PO2 and P6,6,6,14 DEHP dissolved in apolar base oils on alumina surfaces obtained in this study are generally consistent with previous macroscale tribology studies on steel-aluminum surfaces.27 Both studies show that P6,6,6,14 (iC8)2PO2/oil mixtures present obvious breakpoints at longer distances whereas P6,6,6,14 DEHP/oil mixtures do not. These results indicate that the structure and composition of the boundary layers for neat ILs and IL/oil mixtures on alumina are influenced by the composition of ILs. ILs with higher dissociation rates tend to form a more robust boundary layer and thus show higher lubricity.

Conclusion The nanotribology of alumina surfaces lubricated with P6,6,6,14 (iC8)2PO2 and P6,6,6,14 DEHP in hexadecane were investigated using AFM. For both P6,6,6,14 (iC8)2PO2 and P6,6,6,14 DEHP, the neat IL is a much more effective lubricant in comparison to neat hexadecane as seen in previous studies. For IL/oil mixtures, P6,6,6,14 DEHP in hexadecane lubricates the alumina surface better than P6,6,6,14 (iC8)2PO2. Hexadecane is observed to have a strong influence on the integrity of the boundary layer in combination with P6,6,6,14 DEHP or P6,6,6,14 (iC8)2PO2, as the presence of hexadecane assists the replenishment of boundary layers, but reduces the ordering of ions in the boundary layers. For both ILs investigated, the best tribological performance for each IL, is observed for 2 mol% IL in hexadecane, as the boundary layer is robust enough and replenishes quickly. The outcomes of this study help to reveal the effects of IL structure and concentration on lubricity on alumina surfaces, which assists the commercialization of IL lubricants for aluminum and alloys.

14 ACS Paragon Plus Environment

Page 15 of 19 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Supporting Information •

Figure S1. AFM image of Alumina Surface



Figure S2. Repeated friction curves for 0.5 mol% P6,6,6,14 (iC8)2PO2/hexadecane and 0.5 mol% P6,6,6,14 DEHP/hexadecane mixtures



Figure S3. Typical normal force versus apparent separation profile for an AFM tip approaching and retracting from an alumina surface in pure hexadecane (a), pure P6,6,6,14 (iC8)2PO2 (b) and P6,6,6,14 DEHP (c) and 95 mol% P6,6,6,14 DEHP/hexadecane mixture (d)



Figure S4. Lateral force vs normal load for 95% P6,6,6,14 DEHP in Hexadecane.



Table S1. Zero shear viscosity of P6,6,6,14 (iC8)2PO2/hexadecane mixtures at 20 °C measured by an AR-G2 rheometer (TA instruments)



Table S2. Zero shear viscosity of P6,6,6,14 DEHP /hexadecane mixtures at 20 °C measured by an AR-G2 rheometer (TA instruments)

Acknowledgments This research was supported by an Australian Research Council (ARC) Discovery Project (DP120102708). PKC thanks the University of Western Australia for providing a Research Training Program scholarship and AINSE Ltd for providing a Postgraduate Award.

15 ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 16 of 19

References 1. Stachowiak, G. W.; Batchelor, A. W., Engineering Tribology; Butterworth-Heinemann, 2013. 2. Neale, M. J., The Tribology Handbook; Butterworth-Heinemann, 1995. 3. Wan, Y.; Cao, L.; Xue, Q., Friction and Wear Characteristics of Zddp in the Sliding of Steel against Aluminum Alloy. Tribol. Inter. 1997, 30, 767-772. 4. Shimizu, Y.; Spikes, H. A., The Influence of Aluminium–Silicon Alloy on Zddp Tribofilm Formation on the Counter-Surface. Tribol. Lett. 2017, 65, 137. 5. Wasserscheid, P.; Welton, T., Ionic Liquids in Synthesis; Wiley Online Library, 2008; Vol. 1. 6. Liu, W.; Ye, C.; Gong, Q.; Wang, H.; Wang, P., Tribological Performance of Room-Temperature Ionic Liquids as Lubricant. Tribol. Lett. 2002, 13, 81-85. 7. Ye, C.; Liu, W.; Chen, Y.; Yu, L., Room-Temperature Ionic Liquids: A Novel Versatile Lubricant. Chem. Commun. 2001, 2244-2245. 8. Han, Y.; Qiao, D.; Zhang, S.; Feng, D., Influence of Phosphate and Phosphonate Ionic Liquid Structures on Lubrication for Different Alloys (Mg, Al, Cu). Tribol. Inter. 2017, 114, 469-477. 9. Minami, I., Ionic Liquids in Tribology. Molecules 2009, 14, 2286-2305. 10. Somers, A.; Howlett, P.; MacFarlane, D.; Forsyth, M., A Review of Ionic Liquid Lubricants. Lubricants 2013, 1, 3. 11. Somers, A. E.; Biddulph, S. M.; Howlett, P. C.; Sun, J.; MacFarlane, D. R.; Forsyth, M., A Comparison of Phosphorus and Fluorine Containing Il Lubricants for Steel on Aluminium. Phys. Chem. Chem. Phys. 2012, 14, 8224-8231. 12. Somers, A. E.; Howlett, P. C.; Sun, J.; MacFarlane, D. R.; Forsyth, M., Transition in Wear Performance for Ionic Liquid Lubricants under Increasing Load. Tribol. Lett. 2010, 40, 279-284. 13. Perkin, S.; Albrecht, T.; Klein, J., Layering and Shear Properties of an Ionic Liquid, 1-Ethyl-3Methylimidazolium Ethylsulfate, Confined to Nano-Films between Mica Surfaces. Phys. Chem. Chem. Phys. 2010, 12, 1243-1247. 14. Sweeney, J.; Hausen, F.; Hayes, R.; Webber, G. B.; Endres, F.; Rutland, M. W.; Bennewitz, R.; Atkin, R., Control of Nanoscale Friction on Gold in an Ionic Liquid by a Potential-Dependent Ionic Lubricant Layer. Phys. Rev. Lett. 2012, 109, 155502. 15. Carstens, T.; Gustus, R.; Hoefft, O.; Borisenko, N.; Endres, F.; Li, H.; Wood, R. J.; Page, A. J.; Atkin, R., Combined Stm, Afm, and Dft Study of the Highly Ordered Pyrolytic Graphite/1-Octyl-3-MethylImidazolium Bis (Trifluoromethylsulfonyl) Imide Interface. J. .Phys. Chem. C 2014, 118, 10833-10843. 16. Li, H.; Wood, R. J.; Rutland, M. W.; Atkin, R., An Ionic Liquid Lubricant Enables Superlubricity to Be "Switched on" in Situ Using an Electrical Potential. Chem. Comm. 2014, 50, 4368-4370. 17. Elbourne, A.; Sweeney, J.; Webber, G. B.; Wanless, E. J.; Warr, G. G.; Rutland, M. W.; Atkin, R., Adsorbed and near-Surface Structure of Ionic Liquids Determines Nanoscale Friction. Chem. Commun. 2013, 49, 6797-6799. 18. Greaves, T. L.; Weerawardena, A.; Krodkiewska, I.; Drummond, C. J., Protic Ionic Liquids: Physicochemical Properties and Behavior as Amphiphile Self-Assembly Solvents. J. Phys. Chem. B 2008, 112, 896-905. 19. Triolo, A.; Russina, O.; Bleif, H. J.; DiCola, E., Nanoscale Segregation in Room Temperature Ionic Liquids. J. Phys. Chem. B 2007, 111, 4641-4644. 20. Rogers, R. D.; Seddon, K. R., Ionic Liquids--Solvents of the Future? Science 2003, 302, 792-793. 21. Li, H.; Rutland, M. W.; Atkin, R., Ionic Liquid Lubrication: Influence of Ion Structure, Surface Potential and Sliding Velocity. Phys. Chem. Chem. Phys. 2013, 15, 14616-14623.

16 ACS Paragon Plus Environment

Page 17 of 19 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

22. Carstens, T.; Hayes, R.; Abedin, S. Z. E.; Corr, B.; Webber, G. B.; Borisenko, N.; Atkin, R.; Endres, F., In Situ Stm, Afm and Dts Study of the Interface 1-Hexyl-3-Methylimidazolium Tris (Pentafluoroethyl) Trifluorophosphate/Au (111). Electrochim. Acta 2012, 82, 48-59. 23. Li, H.; Wood, R. J.; Endres, F.; Atkin, R., Influence of Alkyl Chain Length and Anion Species on Ionic Liquid Structure at the Graphite Interface as a Function of Applied Potential. J. Phys.: Condens. Matter 2014, 26, 284115. 24. Elbourne, A.; Voïtchovsky, K.; Warr, G. G.; Atkin, R., Ion Structure Controls Ionic Liquid nearSurface and Interfacial Nanostructure. Chem. Sci. 2015, 6, 527-536. 25. Yang, X.; Meng, Y.; Tian, Y., Effect of Imidazolium Ionic Liquid Additives on Lubrication Performance of Propylene Carbonate under Different Electrical Potentials. Tribol. Lett. 2014, 56, 161169. 26. Yu, B.; Bansal, D. G.; Qu, J.; Sun, X.; Luo, H.; Dai, S.; Blau, P. J.; Bunting, B. G.; Mordukhovich, G.; Smolenski, D. J., Oil-Miscible and Non-Corrosive Phosphonium-Based Ionic Liquids as Candidate Lubricant Additives. Wear 2012, 289, 58-64. 27. Somers, A. E.; Khemchandani, B.; Howlett, P. C.; Sun, J.; MacFarlane, D. R.; Forsyth, M., Ionic Liquids as Antiwear Additives in Base Oils: Influence of Structure on Miscibility and Antiwear Performance for Steel on Aluminum. ACS Appl. Mater. Inter. 2013, 5, 11544-11553. 28. Barnhill, W. C.; Qu, J.; Luo, H.; Meyer, H. M.; Ma, C.; Chi, M.; Papke, B. L., PhosphoniumOrganophosphate Ionic Liquids as Lubricant Additives: Effects of Cation Structure on Physicochemical and Tribological Characteristics. ACS Appl. Mater. Inter. 2014, 6, 22585-22593. 29. Qu, J.; Bansal, D. G.; Yu, B.; Howe, J. Y.; Luo, H.; Dai, S.; Li, H.; Blau, P. J.; Bunting, B. G.; Mordukhovich, G., Antiwear Performance and Mechanism of an Oil-Miscible Ionic Liquid as a Lubricant Additive. ACS Appl. Mater. Inter. 2012, 4, 997-1002. 30. Qu, J.; Luo, H.; Chi, M.; Ma, C.; Blau, P. J.; Dai, S.; Viola, M. B., Comparison of an Oil-Miscible Ionic Liquid and Zddp as a Lubricant Anti-Wear Additive. Tribol. Inter. 2014, 71, 88-97. 31. Fan, M.; Yang, D.; Wang, X.; Liu, W.; Fu, H., Doss- Based Qails: As Both Neat Lubricants and Lubricant Additives with Excellent Tribological Properties and Good Detergency. Ind. Eng. Chem. Res. 2014, 53, 17952-17960. 32. Zhou, Y.; Dyck, J.; Graham, T. W.; Luo, H.; Leonard, D. N.; Qu, J., Ionic Liquids Composed of Phosphonium Cations and Organophosphate, Carboxylate, and Sulfonate Anions as Lubricant Antiwear Additives. Langmuir 2014, 30, 13301-13311. 33. Cai, Z.; Meyer, H. M.; Ma, C.; Chi, M.; Luo, H.; Qu, J., Comparison of the Tribological Behavior of Steel–Steel and Si 3 N 4–Steel Contacts in Lubricants with Zddp or Ionic Liquid. Wear 2014, 319, 172-183. 34. Jiménez, A. E.; Bermúdez, M. D.; Carrión, F. J.; Martínez-Nicolás, G., Room Temperature Ionic Liquids as Lubricant Additives in Steel–Aluminium Contacts: Influence of Sliding Velocity, Normal Load and Temperature. Wear 2006, 261, 347-359. 35. Kumeria, T.; Santos, A.; Losic, D., Nanoporous Anodic Alumina Platforms: Engineered Surface Chemistry and Structure for Optical Sensing Applications. Sensors 2014, 14, 11878-11918. 36. Thompson, G., Porous Anodic Alumina: Fabrication, Characterization and Applications. Thin Solid Films 1997, 297, 192-201. 37. Liu, H.; Bhushan, B., Nanotribological Characterization of Molecularly Thick Lubricant Films for Applications to MEMS/NEMS by AFM. Ultramicroscopy 2003, 97, 321-340. 38. Czichos, H., Tribology: A Systems Approach to the Science and Technology of Friction, Lubrication, and Wear; Elsevier, 2009; Vol. 1. 39. Walsh, R. B.; Howard, S. C.; Nelson, A.; Skinner, W. M.; Liu, G.; Craig, V. S. J., Model Surfaces Produced by Atomic Layer Deposition. Chem. Lett. 2012, 41, 1247-1249. 40. Atkin, R.; Warr, G. G., Structure in Confined Room-Temperature Ionic Liquids. J. Phys. Chem. C 2007, 111, 5162-5168.

17 ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 18 of 19

41. Li, H.; Somers, A. E.; Rutland, M. W.; Howlett, P. C.; Atkin, R., Combined Nano- and Macrotribology Studies of Titania Lubrication Using the Oil-Ionic Liquid Mixtures. ACS Sus. Chem. Eng. 2016, 4, 5005-5012. 42. Ralston, J.; Larson, I.; Rutland, M. W.; Feiler, A. A.; Kleijn, M., Atomic Force Microscopy and Direct Surface Force Measurements. Pure Appl. Chem. 2005, 77, 2149-2170. 43. Hadfield, M.; Wang, Y., Tribology in Environmental Design Professional Engineering Publishing Ltd., 2003. 44. Li, H.; Somers, A. E.; Howlett, P. C.; Rutland, M. W.; Forsyth, M.; Atkin, R., Addition of Low Concentrations of an Ionic Liquid to a Base Oil Reduces Friction over Multiple Length Scales: A Combined Nano-and Macrotribology Investigation. Phys. Chem. Chem. Phys. 2016, 18, 6541-6547. 45. Pit, R.; Hervet, H.; Leger, L., Direct Experimental Evidence of Slip in Hexadecane: Solid Interfaces. Phys. Rev. Lett. 2000, 85, 980. 46. Li, H.; Cooper, P. K.; Somers, A. E.; Rutland, M. W.; Howlett, P. C.; Forsyth, M.; Atkin, R., Ionic Liquid Adsorption and Nanotribology at the Silica–Oil Interface: Hundred-Fold Dilution in Oil Lubricates as Effectively as the Pure Ionic Liquid. J. Phys. Chem. Lett. 2014, 5, 4095-4099. 47. Israelachvili, J. N., Forces between Surfaces in Liquids. Adv. Colloid Interf. Sci. 1982, 16, 31-47. 48. Jitvisate, M.; Seddon, J. R. T., Near-Wall Molecular Ordering of Dilute Ionic Liquids. J. Phys. Chem. C 2017, 121, 18593-18597. 49. Battez, A. H.; Bartolomé, M.; Blanco, D.; Viesca, J.; Fernández-González, A.; González, R., Phosphonium Cation-Based Ionic Liquids as Neat Lubricants: Physicochemical and Tribological Performance. Tribol. Inter. 2016, 95, 118-131. 50. Raviv, U.; Klein, J., Fluidity of Bound Hydration Layers. Science 2002, 297, 1540. 51. Jacob, K.; Uri, R.; Susan, P.; Nir, K.; Liraz, C.; Suzanne, G., Fluidity of Water and of Hydrated Ions Confined between Solid Surfaces to Molecularly Thin Films. J. Phys.: Condens. Matter 2004, 16, S5437. 52. Lee, A. A.; Vella, D.; Perkin, S.; Goriely, A., Are Room-Temperature Ionic Liquids Dilute Electrolytes? J. Phys. Chem. Lett. 2015, 6, 159-163.

18 ACS Paragon Plus Environment

Page 19 of 19 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

TOC Graphic

19 ACS Paragon Plus Environment